Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The...

56
Chapter 6: The Fermi Liquid L.D. Landau February 7, 2017 Contents 1 introduction: The Electronic Fermi Liquid 3 2 The Non-Interacting Fermi Gas 5 2.1 Infinite-Square-Well Potential ...................... 5 2.2 The Fermi Gas .............................. 9 2.2.1 T = 0, The Pauli Principle .................... 9 2.2.2 T 6= 0, Fermi Statistics ...................... 12 3 The Weakly Correlated Electronic Liquid 22 3.1 Thomas-Fermi Screening ......................... 22 3.2 The Mott Transition ........................... 25 3.3 Fermi liquids ............................... 28 3.4 Quasi-particles .............................. 29 3.4.1 Particles and Holes ........................ 29 3.4.2 Quasiparticles and Quasiholes at T =0 ............. 33 3.5 Energy of Quasiparticles.......................... 38 4 Interactions between Particles: Landau Fermi Liquid 41 4.1 The free energy, and interparticle interactions ............. 41 1

Transcript of Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The...

Page 1: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

Chapter 6: The Fermi Liquid

L.D. Landau

February 7, 2017

Contents

1 introduction: The Electronic Fermi Liquid 3

2 The Non-Interacting Fermi Gas 5

2.1 Infinite-Square-Well Potential . . . . . . . . . . . . . . . . . . . . . . 5

2.2 The Fermi Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.2.1 T = 0, The Pauli Principle . . . . . . . . . . . . . . . . . . . . 9

2.2.2 T 6= 0, Fermi Statistics . . . . . . . . . . . . . . . . . . . . . . 12

3 The Weakly Correlated Electronic Liquid 22

3.1 Thomas-Fermi Screening . . . . . . . . . . . . . . . . . . . . . . . . . 22

3.2 The Mott Transition . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

3.3 Fermi liquids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3.4 Quasi-particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3.4.1 Particles and Holes . . . . . . . . . . . . . . . . . . . . . . . . 29

3.4.2 Quasiparticles and Quasiholes at T = 0 . . . . . . . . . . . . . 33

3.5 Energy of Quasiparticles. . . . . . . . . . . . . . . . . . . . . . . . . . 38

4 Interactions between Particles: Landau Fermi Liquid 41

4.1 The free energy, and interparticle interactions . . . . . . . . . . . . . 41

1

Page 2: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

4.2 Local Energy of a Quasiparticle . . . . . . . . . . . . . . . . . . . . . 45

4.2.1 Equilibrium Distribution of Quasiparticles at Finite T . . . . . 47

4.2.2 Local Equilibrium Distribution . . . . . . . . . . . . . . . . . 49

4.3 Effective Mass m∗ of Quasiparticles . . . . . . . . . . . . . . . . . . . 51

2

Page 3: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

1 introduction: The Electronic Fermi Liquid

As we have seen, the electronic and lattice degrees of freedom

decouple, to a good approximation, in solids. This is due to the

different time scales involved in these systems.

τion ∼ 1/ωD τelectron ∼h

EF(1)

where EF is the electronic Fermi energy. The electrons may be

thought of as instantly reacting to the (slow) motion of the lat-

tice, while remaining essentially in the electronic ground state.

Thus, to a good approximation the electronic and lattice degrees

of freedom separate, and the small electron-lattice (phonon) in-

teraction (responsible for resistivity, superconductivity etc) may

be treated as a perturbation (with ωD/EF as an expansion pa-

rameter); that is if we are capable of solving the problem of the

remaining purely electronic system.

At first glance the remaining electronic problem would also

appear to be hopeless since the (non-perturbative) electron-

electron interactions are as large as the combined electronic

kinetic energy and the potential energy due to interactions with

3

Page 4: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

the static ions (the latter energy, or rather the corresponding

part of the Hamiltonian, composes the solvable portion of the

problem). However, the Pauli principle keeps low-lying orbitals

from being multiply occupied, so is often justified to ignore the

electron-electron interactions, or treat them as a renormaliza-

tion of the non-interacting problem (effective mass) etc. This

will be the initial assumption of this chapter, in which we will

cover

• the non-interacting Fermi liquid, and

• the renormalized Landau Fermi liquid (Pines & Nozieres).

These relatively simple theories resolve some of the most im-

portant puzzles involving metals at the turn of the century. Per-

haps the most intriguing of these is the metallic specific heat.

Except in certain “heavy fermion” metals, the electronic contri-

bution to the specific heat is always orders of magnitude smaller

than the phonon contribution. However, from the classical theo-

rem of equipartition, if each lattice site contributes just one elec-

tron to the conduction band, one would expect the contributions

from these sources to be similar (Celectron ≈ Cphonon ≈ 3NrkB).

4

Page 5: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

This puzzle is resolved, at the simplest level: that of the non-

interacting Fermi gas.

2 The Non-Interacting Fermi Gas

2.1 Infinite-Square-Well Potential

We will proceed to treat the electronic degrees of freedom, ig-

noring the electron-electron interaction, and even the electron-

lattice interaction. In general, the electronic degrees of freedom

are split into electrons which are bound to their atomic cores

with wavefunctions which are essentially atomic, unaffected by

the lattice, and those valence (or near valence) electrons which

react and adapt to their environment. For the most part, we are

only interested in the valence electrons. Their environment de-

scribed by the potential due to the ions and the core electrons–

the core potential. Thus, ignoring the electron-electron interac-

tions, the electronic Hamiltonian is

H =P 2

2m+ V (r) . (2)

5

Page 6: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

As shown in Fig. 1, the core potential V (r), like the lattice, is

periodic

a

V(r)

V(r+a) = V(r)

Figure 1: Schematic core potential (solid line) for a one-dimensional lattice with

lattice constant a.

For the moment, ignore the core potential, then the electronic

wave functions are plane waves ψ ∼ eik·r. Now consider the

core potential as a perturbation. The electrons will be strongly

effected by the periodicity of the potential when λ = 2π/k ∼ a

1. However, when k is small so that λ a (or when k is large,

so λ a) the structure of the potential may be neglected, or

we can assume V (r) = V0 anywhere within the material. The

potential still acts to confine the electrons (and so maintain

charge neutrality), so V (r) =∞ anywhere outside the material.

1Interestingly, when λ ∼ a, the Bragg condition 2d sin θ ≈ a ≈ λ may easily be satisfied, so the

electrons, which may be though of as DeBroglie waves, scatter off of the lattice. Consequently states

for which λ = 2π/k ∼ a are often forbidden. This is the source of gaps in the band structure, to be

discussed in the next chapter.

6

Page 7: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

Figure 2: Infinite square-well potential. V (r) = V0 within the well, and V (r) = ∞outside to confine the electrons and maintain charge neutrality.

Thus we will approximate the potential of a cubic solid with

linear dimension L as an infinite square-well potential.

V (r) =

V0 0 < ri < L

∞ otherwise(3)

The electronic wavefunctions in this potential satisfy

− h2

2m∇2ψ(r) = (E ′ − V0)ψ(r) = Eψ(r) (4)

The normalize plane wave solution to this model is

ψ(r) =

3∏i=1

(2

L

)1/2

sin kixi where i = x, y, or z (5)

and kiL = niπ in order to satisfy the boundary condition that

ψ = 0 on the surface of the cube. Furthermore, solutions with

ni < 0 are not independent of solutions with ni > 0 and may

7

Page 8: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

be excluded. Solutions with ni = 0 cannot be normalized and

are excluded (they correspond to no electron in the state). The

π/L

kx

ky

kz

(π/L)3

Figure 3: Allowed k-states for an electron confined by a infinite-square potential. Each

state has a volume of (π/L)3 in k-space.

eigenenergies of the wavefunctions are

−h2∇2

2mψ =

h2

2m

∑i

k2i =

h2π2

2mL2

(n2x + n2

y + n2z

)(6)

and as a result of these restrictions, states in k-space are con-

fined to the first quadrant (c.f. Fig. 3). Each state has a volume

(π/L)3 of k-space. Thus as L→∞, the number of states with

energies E(k) < E < E(k) + dE is

dZ ′ =(4πk2dk)/8

(π/L)3. (7)

8

Page 9: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

Then, since E = h2k2

2m , so k2dk = mh2

√2mE/h2dE

dZ = dZ ′/L3 =1

4π2

(2m

h2

)3/2

E1/2dE . (8)

or, the density of state per unit volume is

D(E) =dZ

dE=

1

4π2

(2m

h2

)3/2

E1/2 . (9)

Up until now, we have ignored the properties of electrons. How-

ever, for the DOS, it is useful to recall that the electrons are

spin-1/2 thus 2S + 1 = 2 electrons can fill each orbital or k-

state, one of spin up the other spin down. If we account for this

spin degeneracy in D, then

D(E) =1

2π2

(2m

h2

)3/2

E1/2 . (10)

2.2 The Fermi Gas

2.2.1 T = 0, The Pauli Principle

Electrons, as are all half-integer spin particles, are Fermions.

Thus, by the Pauli Principle, no two of them may occupy the

same state. For example, if we calculate the density of electrons

9

Page 10: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

per unit volume

n =

∫ ∞0

D(E)f (E, T )dE , (11)

where f (E, T ) is the probability that a state of energy E is oc-

cupied, the factor f (E, T ) must enforce this restriction. How-

ever, f is just the statistical factor; c.f. for classical particles

f (E, T ) = e−E/kBT for classical particles , (12)

which for T = 0 would require all the electrons to go into

the ground state f (0, 0) = 1. Clearly, this violates the Pauli

principle.

At T = 0 we need to put just one particle in each state, start-

ing from the lowest energy state, until we are out of particles.

Since E ∝ k2 in our simple square-well model, will fill up all

k-states until we reach some Fermi radius kF , corresponding to

some Fermi Energy EF

EF =h2k2

F

2m, (13)

thus,

f (E, T = 0) = θ(EF − E) (14)

10

Page 11: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

kx

ky

kz

k f

k f

occupied states

h k 2 2

f2m

= Ef

D(E)

Figure 4: Due to the Pauli principle, all k-states up to kF , and all states with energies

up to Ef are filled at zero temperature.

and

n =

∫ ∞0

D(E)f (E, T )dE =

∫ EF

0

D(E)DE

=

(2m

h2

)3/21

2π2

∫ EF

0

E1/2DE

=

(2m

h2

)3/21

2π2

2

3E

3/2F , (15)

or

EF =h2

2m

(3π2n

)2/3= kBTF (16)

which also defines the Fermi temperature TF . Thus for metals,

in which n ≈ 1023/cm3, EF ≈ 10−11erg ≈ 10eV ≈ kB105K.

Notice that due to the Pauli principle, the average energy of

11

Page 12: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

the electrons will be finite, even at T = 0!

E =

∫ EF

0

D(E)EdE =3

5nEF . (17)

However, it is the electrons near EF in energy which may be

excited and are therefore important. These have a DeBroglie

wavelength of roughly

λe =12.3

A

(E(eV))1/2≈ 4

A (18)

thus our original approximation of a square well potential, ig-

noring the lattice structure, is questionable for electrons near

the Fermi surface, and should be regarded as yielding only qual-

itative results.

2.2.2 T 6= 0, Fermi Statistics

At finite temperatures some of the states will be thermally ex-

cited. The energy available for these excitations is roughly

kBT , and the only possible excitations are from filled to un-

filled electronic states. Therefore, only the states within kBT

(EF − kBT < E < EF + kBT ) of the Fermi surface may be

excited. f (E, T ) must be modified accordingly.

12

Page 13: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

What we need is then f (E, T ) at finite T which also satisfies

the Pauli principle. Lets return to our model of a periodic solid

which is constructed by bringing individual atoms together from

an infinite separation. First, just consider a solid constructed

from only two atoms, each with a single orbital (Fig. 5). For

1 2

δ n = -11

δ n = +11

Figure 5: Exchange of electrons in a solid composed of two orbitals.

this system, in equilibrium,

0 = δF =∑i

∂F

∂niδni (19)

electrons are conserved so∑

i δni = 0. Thus, for our two orbital

system

∂F

∂n1δn1 +

∂F

∂n2δn2 = 0 and δn1 + δn2 = 0 (20)

or∂F

∂n1=∂F

∂n2(21)

13

Page 14: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

A similar relation holds for an arbitrary number of particles.

Apparently this quantity, the increased free energy needed to

add a particle to the system, is a constant

∂F

∂ni= µ (22)

for all i. µ is called the chemical potential.

Now consider an ensemble of orbitals. We will treat the ther-

modynamics of this system within the canonical ensemble (i.e.,

the system is in contact with a thermal bath, and the particle

number is conserved) for which F = E −TS is the appropriate

potential. The system energy E and Entropy S may be written

as functions of the orbital energies Ei and occupancies ni and

the degeneracy gi of the state of energy Ei. For example,

g = 21E1

E2

E3 g = 43

g = 22

E4 g = 44

n = 21

n = 43

n = 12

n = 24

Figure 6: states from an ensemble of orbitals.

E =∑i

niEi . (23)

14

Page 15: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

The entropy S requires a bit more thought. If P is the number

of ways of distributing the electrons among the states, then

S = kB lnP . (24)

Consider a set of gi states with energy Ei. The number of

ways of distributing the first electron in these states is gi. For

a second electron we then have gi − 1 ways... etc. So for ni

electrons there aregi!

ni!(gi − ni)!(25)

possible ways of accommodating the ni (indistinguishable) elec-

trons in gi states.

The number of ways of making the whole system (ie, filling

energy levels with Ei 6= Ej) is then

P =∏i

gi!

ni!(gi − ni)!, (26)

and so, the entropy

S = kB∑i

ln gi!− lnni!− ln(gi − ni)! . (27)

For large n, lnn! ≈ n lnn− n, so

S = kB∑i

gi ln gi − ni lnni − (gi − ni) ln(gi − ni) (28)

15

Page 16: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

and

F =∑i

niEi−kBT∑i

gi ln gi−ni lnni− (gi−ni) ln(gi−ni)

(29)

We will want to use the chemical potential µ in our thermo-

dynamic calculations

µ =∂F∂nk

= Ek + kBT (lnnk + 1− ln(gk − nk)− 1) , (30)

where β = 1/kBT . Solving for nk

nk =gk

1 + eβ(Ek−µ). (31)

Thus the probability that a quantum state with energy E is

occupied, is (the Fermi function)

f (E, T ) =1

1 + eβ(Ek−µ). (32)

At T = 0, β =∞, and f (E, 0) = θ(µ−E). Thus µ(T = 0) =

EF . However in general µ is temperature dependent, since it

must be adjusted to keep the particle number fixed. In addition,

when T 6= 0, f becomes less sharp at energies E ≈ µ. This

reflects the fact that particles with energies E −µ ≈ kBT may

be excited to higher energy states.

16

Page 17: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

0.0 0.5 1.0 1.5 2.0ω

0.0

0.5

1.0

1.5

1/(e

β(ω

−µ) +

1)

Figure 7: Plot of the Fermi function 1/(e−β(ω−µ) + 1

)when β = 1/kBT = 20 and

µ = 1. Not that at energies ω ≈ µ the Fermi function displays a smooth step of width

≈ kBT = 0.05. This allows thermal excitations of particles near the Fermi surface.

Specific Heat The form of f (E, T ) also clarifies why the elec-

tronic specific heat of metals is so small compared to the clas-

sical result Cclassical = 32nkBT . The reason is simple: only the

electrons with energies within about kBT of the Fermi surface

may be excited (about kBTEF

of the electron density) each with

excitation energy of about kBT . Therefore,

Uexcitation ≈ kBTnkBT

EF= nkBT

T

TF(33)

so

C ≈ nkBT

TF(34)

17

Page 18: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

Then as T TF (TF is typically about 105K in most metals2)

C ≈ nkBTTF Cclassical ≈ nkB. Thus at temperatures where

the phonons contribute essentially a classical result to the spe-

cific heat, the electronic contribution is vanishingly small. In

general this holds except at very low T where the phonon con-

tribution Cphonon ∼ T 3 goes to zero faster than the electronic

contribution to the specific heat.

Specific Heat Calculation Of course, since we know the free energy

of the non-interacting Fermi gas, we can calculate the form of

the specific heat. Here we will follow Ibach and Luth and Kittel;

however, since the chemical potential does depend upon the

temperature, I would like to make the approximations we make

a bit more explicit.

Upon heating from T = 0 to finite T , the Fermi gas will gain

energy

U(T ) =

∫ ∞0

dE ED(E)f (E, T )−∫ EF

0

dE ED(E) (35)

2Heavy Fermion systems are the exception to this rule. There TF can be as small as a fraction

of a degree Kelvin. As a result, they may have very large electronic specific heats.

18

Page 19: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

so

CV =dU

dT=

∫ ∞0

dE ED(E)df (E, T )

dT. (36)

Then since at constant volume the electronic density is constant,

so dndT = 0, and n =

∫∞0 dED(E)f (E, T ),

0 = EFdn

dT=

∫ ∞0

dE EFD(E)df (E, T )

dT(37)

so we may write

CV =

∫ ∞0

dE (E − EF )D(E)df

dT. (38)

In f , the temperature T enters through both β = 1/kBT and

µ

df

dT=

∂f

∂β

∂β

∂T+∂f

∂µ

∂µ

∂T

=βeβ(E−µ)(

eβ(E−µ) + 1)2

[E − µT− ∂µ

∂T

](39)

However, ∂µ∂T depends upon the details of the density of states

near the Fermi surface, which can differ greatly from material to

material. Furthermore,∣∣∣ ∂µ∂T ∣∣∣ < 1 especially in common metals

at temperatures T TF , and the first term E−µT is of order

one (c.f. Fig. 7). Thus, for now we will neglect ∂µ∂T relative to

E−µT (you will explore the validity of this approximation in your

19

Page 20: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

homework), and, consistent with this approximation, replace µ

by EF , sodf

dT≈ βeβ(E−EF )(

eβ(E−EF ) + 1)2

E − EF

T, (40)

and

CV ≈1

kBT 2

∫ ∞0

dE(E − EF )2D(E)βeβ(E−EF )(

eβ(E−EF ) + 1)2 let x = β(E − EF )

≈ kBT

∫ ∞−βEF

dxD

(x

β+ EF

)x2 ex

(ex + 1)2 (41)

As shown in Fig. 8, the function x2 ex

(ex+1)2 is only large in the

region −10 < x < 10. In this region, and for temperatures

-10 -5 0 5 10x

0.0

0.1

0.2

0.3

0.4

0.5

x2 ex /(ex +

1)2

Figure 8: Plot of x2 ex

(ex+1)2vs. x. Note that this function is only finite for roughly

−10 < x < 10. Thus, at temperatures T TF ∼ 105K, we can approximate

D(xβ

+ EF

)≈ D (EF ) in Eq. 42.

T TF , D(xβ + EF

)≈ D (EF ), since the density of states

20

Page 21: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

usually does not have features which are sharp on the energy

scale of 10kBT . Thus

CV ≈ kBTD (EF )

∫ ∞−βEF

dx x2 ex

(ex + 1)2

≈ π2

3k2BTD(EF ) . (42)

Note that no assumption about the form of D(E) was made

other than the assumption that it is smooth within kBT of the

Fermi surface. Thus, experimental measurements of the specific

heat at constant volume of the electrons, gives us information

about the density of electronic states at the Fermi surface.

Now let’s reconsider the DOS for the 3-D box potential.

D(E) =1

2π2

(2m

h

)3/2

E1/2 = D(EF )

(E

EF

)1/2

(43)

For which n =∫ EF

0 D(E)dE = D(EF )23EF , so

CV =π2

2nkB

T

TF 3

2nkB (44)

where the last term on the right is the classical result. For

room temperatures T ∼ 300K, which is also of the same order

of magnitude as the Debye temperatures θD,

CV phonon ∼3

2nkB CV electron (45)

21

Page 22: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

So, the only way to measure the electronic specific heat in most

materials is to go to very low temperatures T θD, for which

CV phonon ∼ T 3. Here the total specific heat

CV ≈ γT + βT 3 (46)

We will see that gives us some measurement of the electronic ef-

fective mass for our Fermi liquid theory. I.e. it tell us something

about electron- electron interactions.

3 The Weakly Correlated Electronic Liquid

3.1 Thomas-Fermi Screening

As an introduction to the effect of electronic correlations, con-

sider the effect of a charged oxygen defect in one of the copper-

oxygen planes of a cuprate superconductor shown in Fig. 9.

Assume that the oxygen defect captures two electrons from the

metallic band, going from a 2s22p4 to a 2s22p6 configuration.

The defect will then become a cation, and have a net charge

of two electrons. In the vicinity of this oxygen defect, the elec-

trostatic potential and the electronic charge density will be re-

22

Page 23: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

Cu O

O

Cu O

O

Cu O

O

Cu O

O

Cu O

O

Cu O

O

Cu O

O

Cu O

O

Cu O

O

O

Cu O

O

Cu O

O

Cu O

O

Cu O

O

Cu O

O

Cu O

O

Cu O

O

Cu O

O

Cu O

O

Oq=2e-

Figure 9: A charged oxygen defect is introduced into one of the copper-oxygen planes of

a cuprate superconductor. The oxygen defect captures two electrons from the metallic

band, going from a 2s22p4 to a 2s22p6 configuration.

duced.

If we model the electronic density of states in this material

with our box-potential DOS, we can think of this reduction in

the local charge density in terms of raising the DOS parabola

near the defect (cf. Fig. 10). This will cause the free electronic

charge to flow away from the defect. Near the defect (since

e < 0 and hence eδU(rnear) < 0)

n(rnear) ≈∫ EF+eδU(rnear)

0

D(E)DE (47)

While away from the defect, δU(raway) = 0, so

n(raway) ≈∫ EF

0

D(E)DE (48)

23

Page 24: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

-eδU

EF

e

near chargeddefect

Away fromcharged defect

Figure 10: The shift in the DOS parabola near a charged defect.

or

δn(r) ≈∫ EF+eδU(r)

0

D(E)DE −∫ EF

0

D(E)DE (49)

If |eδU | EF , then

δn(r) ≈ eδUD(EF ) . (50)

We can solve for the change in the electrostatic potential by

solving Poisson equation.

∇2δU = 4πδρ = 4πeδn = 4πe2D(EF )δU . (51)

Let λ2 = 4πe2D(EF ), then ∇2δU = λ2δU has the solution3

δU(r) =qe−λr

r(52)

3The solution is actually Ce−λr/r, where C is a constant. C may be deterined by letting D(EF ) =

0, so the medium in which the charge is embedded becomes vacuum. Then the potential of the charge

is q/r, so C = q.

24

Page 25: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

The length 1/λ = rTF is known as the Thomas-Fermi screening

length.

rTF =(4πe2D(EF )

)−1/2(53)

Lets estimate this distance for our square-well model,

r2TF =

a0π

3(3π2n)1/3≈ a0

4n1/3

rTF ≈1

2

(n

a30

)−1/6

(54)

In Cu, for which n ≈ 1023 cm−3 (and since a0 = 0.53A)

rTFCu ≈1

2

(1023

)−1/6

(0.5× 10−8)−1/2≈ 0.5× 10−8 cm = 0.5

A (55)

Thus, if we add a charge defect to Cu metal, the effect of the

defect’s ionic potential is screened away for distances r > 12

A!

3.2 The Mott Transition

Now consider an electron bound to an ion in Cu or some other

metal. As shown in Fig. 11 the screening length decreases, and

bound states rise up in energy. In a weak metal (i.e., something

like YBCO), in which the valence state is barely free, a reduction

in the number of carriers (electrons) will increase the screening

25

Page 26: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

-e-r

/rT

F /r

rTF=1/4

r

rTF=1

r

-e-r

/rT

F /r

bound statesfree states

rTF= n-1/6

Figure 11: Screened defect potentials. As the screening length increases, states that

were free, become bound.

length, since

rTF ∼ n−1/6 . (56)

This will extend the range of the potential, causing it to trap

or bind more states–making the one free valance state bound.

Now imagine that instead of a single defect, we have a con-

centrated system of such ions, and suppose that we decrease the

density of carriers (i.e., in Si-based semiconductors, this is done

by doping certain compensating dopants, or even by modulating

the pressure). This will in turn, increase the screening length,

causing some states that were free to become bound, causing an

abrupt transition from a metal to an insulator, and is believed

to explain the MI transition in some transition-metal oxides,

glasses, amorphous semiconductors, etc.This metal-insulator tran-

sition was first proposed by N. Mott, and is called the Mott

26

Page 27: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

transition, more significantly Mott proposed a criterion based

on the relevant electronic density for when this transition should

occur.

In Mott’s criterion, a metal-insulator transition occurs when

this potential, in this case coming from the addition of an ionic

impurity, can just bind an electronic state. If the state is bound,

the impurity band is localized. If the state is not bound, then

the impurity band is extended. The critical value of λ = λc

may be determined numerically[1] with λc/a0 = 1.19, which

yields the Mott criterion of

2.8a0 ≈ n−1/3c , (57)

where a0 is the Bohr radius. Despite the fact that electronic in-

teractions are only incorporated in the extremely weak coupling

limit, Thomas-Fermi Screening, Mott’s criterion even works for

moderately and strongly interacting systems[2].

27

Page 28: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

3.3 Fermi liquids

The purpose of these next several lectures is to introduce you

to the theory of the Fermi liquid, which is, in its simplest form,

a collection of Fermions in a box plus interactions.

In reality , the only physical analog is a gas of 3He, which

due its nuclear spin (the nucleus has two protons, one neutron),

obeys Fermi statistics for sufficiently low energies or temper-

atures. In addition, simple metals, from the first or second

column of the periodic table, for which we may approximate

the ionic potential

V (R) = V0 (58)

are a close approximant to Fermi liquids.

Moreover, Fermi Liquid theory only describes the ”gaseous”

phase of these quantum fermion systems. For example, 3He also

has a superfluid (triplet), and at least in 4He-3He mixtures, a

solid phase exists which is not described by Fermi Liquid The-

ory. One should note; however, that the Fermi liquid theory

state does serve as the starting point for the theories of super-

conductivity and super fluidity.

28

Page 29: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

One may construct Fermi liquid theory either starting from a

many-body diagrammatic or phenomenological viewpoint. We,

as Landau, will choose the latter. Fermi liquid theory has 3

basic tenants:

1. momentum and spin remain good quantum numbers to de-

scribe the (quasi) particles.

2. the interacting system may be obtained by adiabatically

turning on a particle-particle interaction over some time t.

3. the resulting excitations may be described as quasi-particles

with lifetimes t.

3.4 Quasi-particles

The last assumption involves a new concept, that of the quasi-

particles which requires some explanation.

3.4.1 Particles and Holes

Particles and Holes are excitations of the non-interacting system

at zero temperature. Consider a system of N free Fermions

29

Page 30: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

each of mass m in a volume V . The eigenstates are the anti-

symmetrized combinations (Slater determinants) of N different

single particle states.

ψp(r) =1√Veip·r/h (59)

The occupation of each of these states is given by np = θ(p−pF )

where pF is the radius of the Fermi sphere. The energy of the

system is

E =∑

p

npp2

2m(60)

and pF is given by

N

V=

1

3π2

(pFh

)3

(61)

Now lets add a particle to the lowest available state p = pF

then, for T = 0,

µ = E0(N + 1)− E0(N) =∂E0

∂N=p2F

2m. (62)

If we now excite the system, we will promote a certain number

of particles across the Fermi surface SF yielding particles above

and an equal number of vacancies or holes below the Fermi

surface. These are our elementary excitations, and they are

30

Page 31: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

quantified by δnp = np − n0p

δnp =

δp,p′ for a particle p′ > pF

−δp,p′ for a hole p′ < pF. (63)

If we consider excitations created by thermal fluctuations, then

EF

E

D(E)

particleexcitation

hole excitation

δn = -1p

δn = 1p’

Figure 12: Particle and hole excitations of the Fermi gas.

δnp ∼ 1 only for excitations of energy within kBT of EF . The

energy of the non-interacting system is completely characterized

as a functional of the occupation

E − E0 =∑

p

p2

2m(np − n0

p) =∑

p

p2

2mδnp . (64)

Now lets take our system and place it in contact with a par-

ticle bath. Then the appropriate potential is the free energy,

31

Page 32: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

E

D(E)

δF = p’ /2m - µ2

δF = µ - p /2m 2

E = µ = p /2mF 2

F

δF = | µ - p /2m | 2

Figure 13: Since µ = p2F/2m, the free energy of a particle or a hole is δF =

|p2/2m− µ| > 0, so the system is stable to these excitations.

which for T = 0, is F = E − µN , and

F − F0 =∑p

(p2

2m− µ

)δnp . (65)

The free energy of a particle, with momentum p and δnp′ =

δp,p′ is p2

2m − µ and it corresponds to an excitation outside SF .

The free energy of a hole δnp′ = −δp,p′ is µ− p2

2m, which corre-

sponds to an excitation within SF . However, since µ = p2F/2m,

the free energy of either at p = pF is zero, hence the free energy

of an excitation is ∣∣p2/2m− µ∣∣ , (66)

which is always positive; ie., the system is stable to excitations.

32

Page 33: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

aU ≈ ee

2

a -a/r TF

Figure 14: Model for a fermi liquid: a set of interacting particles an average distance

a apart bound within an infinite square-well potential.

3.4.2 Quasiparticles and Quasiholes at T = 0

Now let’s consider a system with interacting particles an aver-

age distance a apart, so that the characteristic energy of inter-

action is e2

a e−a/rTF . We will imagine that this system evolves

slowly from an ideal or noninteracting system in time t (i.e.,

the interaction U ≈ e2

a e−a/rTF is turned on slowly, so that the

non-interacting system evolves while remaining in the ground

state into an interacting system in time t).

If the eigenstate of the ideal system is characterized by n0p,

then the interacting system eigenstate will evolve quasistatis-

tically from n0p to np. In fact if the system is isotropic and

remains in its ground state, then n0p = np. However, clearly in

some situations (superconductivity, magnetism) we will neglect

33

Page 34: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

some eigenstates of the interacting system in this way.

Now let’s add a particle of momentum p to the non-interacting

ideal system, and slowly turn on the interaction. As U is

p

time = 0U = 0

p p

time = tU = (e /a) exp(-a/r )

2TF

Figure 15: We add a particle with momentum p to our noninteracting (U = 0) Fermi

liquid at time t = 0, and slowly increase the interaction to its full value U at time t.

As the particle and system evolve, the particle becomes dressed by interactions with

the system (shown as a shaded ellipse) which changes the effective mass but not the

momentum of this single-particle excitation (now called a quasi-particle).

switched on, we slowly begin to perturb the particles close to

the additional particle, so the particle becomes dressed by these

interactions. However since momentum is conserved, we have

created an excitation (particle and its cloud) of momentum p.

We call this particle and cloud a quasiparticle. In the same way,

if we had introduced a hole of momentum p below the Fermi

surface, and slowly turned on the interaction, we would have

produced a quasihole.

Note that this adiabatic switching on procedure will have

34

Page 35: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

difficulties if the lifetime of the quasi-particle τ < t. If so,

then the the quasiparticle could decay into a number of other

quasiparticles and quasiholes. If we shorten t so that again

τ t, then switching on U could excite the system out of its

ground state). We will show that such difficulties do not arise so

long as the energy of the particle is close to the Fermi energy.

Here there are few states accessible for creating particle-hole

excitations so collisions are rare.

p1

p4

p2

p3

Figure 16: A particle with momentum p1 above the Fermi surface (p1 > pF ) interacts

with one of the particles below the Fermi surface with momentum p2. As a result, two

new particles appear above the Fermi surface (all other states are full) with momenta

p3 and p4..

To estimate this lifetime consider the following argument

from AGD: A particle with momentum p1 above the Fermi

35

Page 36: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

surface (p1 > pF ) interacts with one of the particles below the

Fermi surface with momentum p2. As a result, two new par-

ticles appear above the Fermi surface (all other states are full)

with momenta p3 and p4. This may also be interpreted as a

particle of momentum p1 decaying into particles with momenta

p3 and p4 and a hole with momentum p2. By Fermi’s golden

rule, the total probability of such a process if proportional to

1

τ∝∫δ (ε1 + ε2 − ε3 − ε4) d3p2d

3p3 (67)

where ε1 =p2

12m − EF , and the integral is subject to the con-

straints of energy and momentum conservation and that

p2 < pF , p3 > pF , p4 = |p1 + p2 − p3| > pF (68)

It must be that ε1 + ε2 = ε3 + ε4 > 0 since both particles 3

and 4 must be above the Fermi surface. However, since ε2 < 0,

if ε1 is small, then |ε2| <∼ ε1 is also small, so only of order

ε1/EF states may scatter with the state k1, conserve energy,

and obey the Pauli principle. Thus, restricting ε2 to a narrow

shell of width ε1/EF near the Fermi surface, and reducing the

scattering probability 1/τ by the same factor.

36

Page 37: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

k1 k

4

EF3

2 1

4

k - k1 3

k - k4 2

E

N(E)

ky

kx

k3

k2

Figure 17: A quasiparticle of momentum p1 decays via a particle-hole excitation

into a quasiparticle of momentum p4. This may also be interpreted as a particle

of momentum p1 decaying into particles with momenta p3 and p4 and a hole with

momentum p2. Energy conservation requires |ε2| <∼ ε1. Thus, restricting ε2 to a

narrow shell of width ε1/EF near the Fermi surface. Momentum conservation k1 −k3 = k4 − k2 further restricts the available states by a factor of about ε1/EF . Thus

the lifetime of a quasiparticle is proportional to(ε1EF

)−2

.

Now consider the constraints placed on states k3 and k4 by

momentum conservation

k1 − k3 = k4 − k2 . (69)

Since ε1 and ε2 are confined to a narrow shell around the Fermi

surface, so too are ε3 and ε4. This can be seen in Fig. 17, where

the requirement that k1 − k3 = k4 − k2 limits the allowed

states for particles 3 and 4. If we take k1 fixed, then the allowed

states for 2 and 3 are obtained by rotating the vectors k1−k3 =

k4−k2; however, this rotation is severely limited by the fact that

37

Page 38: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

particle 3 must remain above, and particle 2 below, the Fermi

surface. This restriction on the final states further reduces the

scattering probability by a factor of ε1/EF .

Thus, the scattering rate 1/τ is proportional to(ε1EF

)2

so

that excitations of sufficiently small energy will always be suf-

ficiently long lived to satisfy the constraints of reversibility. Fi-

nally, the fact that the quasiparticle only interacts with a small

number of other particles due to Thomas-Fermi screening (i.e.,

those within a distance ≈ rTF ), also significantly reduces the

scattering rate.

3.5 Energy of Quasiparticles.

As in the non-interacting system, excitations will be quanti-

fied by the deviation of the occupation from the ground state

occupation n0p

δnp = np − n0p . (70)

At low temperatures δnp ∼ 1 only for p ≈ pF where the par-

ticles are sufficiently long lived that τ t. It is important to

emphasize that only δnp not n0p or np, will be physically rele-

38

Page 39: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

vant. This is important since it does not make much sense to

talk about quasiparticle states, described by np, far from the

Fermi surface since they are not stable and cannot be excited

by the perturbations we are considering (i.e., thermal, or in a

transport experiment).

For the ideal system

E − E0 =∑

p

p2

2mδnp . (71)

For the interacting system E[np] becomes much more compli-

cated. If however δnp is small (so that the system is close to

its ground state) then we may expand:

E[np] = Eo +∑

p

εpδnp +O(δn2p) , (72)

where εp = δE/δnp. Note that εp is intensive (ie. it is indepen-

dent of the system volume). If δnp = δp,p′, then E ≈ E0 + εp′;

i.e., the energy of the quasiparticle of momentum p′ is εp′.

In practice we will only need εp near the Fermi surface where

δnp is finite. So we may approximate

εp ≈ µ + (p− pF ) · ∇p εp|pF (73)

39

Page 40: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

where ∇pεp = vp, the group velocity of the quasiparticle. The

ground state of the N +1 particle system is obtained by adding

a particle with εp = εF = µ = ∂E0∂N (at zero temperature); which

defines the chemical potential µ. We make learn more about

εp by employing the symmetries of our system. If we explicitly

display the spin-dependence,

εp,σ = ε−p,−σ under time-reversal (74)

εp,σ = ε−p,σ under BZ reflection (75)

So εp,σ = ε−p,σ = εp,−σ; i.e., in the absence of an external

magnetic field, εp,σ does not depend upon σ if. Furthermore,

for an isotropic system εp depends only upon the magnitude of

p, |p|, so p and vp = ∇εp(|p|) = p|p|dεp(|p|)d|p| are parallel. Let us

define m∗ as the constant of proportionality at the fermi surface

vpF = pF/m∗ (76)

Using m∗ it is useful to define the density of states at the

fermi surface. Recall, that in the non-interacting system,

D(EF ) =1

2π2

(2m

h2

)3/2

E1/2F =

mpF

πh3 (77)

40

Page 41: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

where p = hk, and E = p2/2m. Thus, for the interacting

system at the Fermi surface

Dinteracting(EF ) =m∗pF

πh3 , (78)

where the m∗ (generally > m, but not always) accounts for the

fact that the quasiparticle may be viewed as a dressed particle,

and must “drag” this dressing along with it. I.e., the effective

mass to some extent accounts for the interaction between the

particles.

4 Interactions between Particles: Landau Fermi

Liquid

4.1 The free energy, and interparticle interactions

The thermodynamics of the system depends upon the free en-

ergy F , which at zero temperature is

F − F0 = E − E0 − µ(N −N0) . (79)

Since our quasiparticles are formed by adiabatically switching

on the interaction in the N + 1 particle ideal system, adding

41

Page 42: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

one quasiparticle to the system adds one real particle. Thus,

N −N0 =∑

p

δnp , (80)

and since

E − E0 ≈∑

p

εpδnp , (81)

we get

F − F0 ≈∑

p

(εp − µ) δnp . (82)

As shown in Fig. 18, we will be interested in excitations of the

system which distort the Fermi surface by an amount propor-

tional to δ. For our theory/expansion to remain valid, we must

δ

Figure 18: We consider small distortions of the fermi surface, proportional to δ, so

that 1N

∑p |δnp| 1.

42

Page 43: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

have1

N

∑p

|δnp| 1 . (83)

Where δnp 6= 0, εp − µ will also be of order δ. Thus,∑p

(εp − µ) δnp ∼ O(δ2) , (84)

so, to be consistent we must add the next term in the Taylor

series expansion of the energy to the expression for the free

energy.

F −F0 =∑

p

(εp − µ) δnp +1

2

∑p,p′

fp,p′δnpδnp′+O(δ3) (85)

where

fp,p′ =δE

δnpδnp′(86)

The term, proportional to fp,p′, was added (to the Sommerfeld

theory) by L.D. Landau. Since each sum over p is proportional

to the volume V , as is F , it must be that fp,p′ ∼ 1/V . However,

it is also clear that fp,p′ is an interaction between quasiparticles,

each of which is spread out over the whole volume V , so the

probability that they will interact is ∼ r3TF/V , thus

fp,p′ ∼ r3TF/V

2 (87)

43

Page 44: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

In general, since δnp is only of order one near the Fermi

surface, we will only care about fp,p′ on the Fermi surface (as-

suming that it is continuous and changes slowly as we cross the

Fermi surface.

Interested only in fp,p′∣∣εp=εp′=µ

! (88)

Thus, fp,p′ only depends upon the angle between p and p′.

We can also reduce the spin dependence of fp,p′ to a symmet-

ric and anti symmetric part. First in the absence of an external

field, the system should be invariant under time-reversal, so

fpσ,p′σ′ = f−p−σ,−p′−σ′ , (89)

and, in a system with reflection symmetry

fpσ,p′σ′ = f−pσ,−p′σ′ . (90)

Then

fpσ,p′σ′ = fp−σ,p′−σ′ . (91)

It must be then that f depends only upon the relative orienta-

tions of the spins σ and σ′, so there are only two independent

components fp↑,p′↑ and fp↑,p′↓. We can split these into sym-

44

Page 45: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

metric and antisymmetric parts.

fap,p′ =1

2

(fp↑,p′↑ − fp↑,p′↓

)f sp,p′ =

1

2

(fp↑,p′↑ + fp↑,p′↓

).

(92)

fap,p′ may be interpreted as an exchange interaction, or

fpσ,p′σ′ = f sp,p′ + σ · σ′fap,p′ (93)

where σ and σ′ are the Pauli matrices for the spins.

Our ideal system is isotropic in momentum. Thus, fap,p′ and

f sp,p′ will only depend upon the angle θ between p and p′, and

so we may expand either fap,p′ and f sp,p′

fαp,p′ =

∞∑l=0

fαl Pl(cos θ) . (94)

Conventionally these f parameters are expressed in terms of

reduced units.

D(EF )fαl =V m∗pF

π2h3 fαl = F αl . (95)

4.2 Local Energy of a Quasiparticle

Now consider an interacting system with a certain distribution

of excited quasiparticles δnp′. To this, add another quasiparticle

45

Page 46: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

p

Figure 19: The addition of another particle to a homogeneous system will yeilds in

forces on the quasiparticle which tend to restore equilibrium.

of momentum p (δn′p → δn′p + δp,p′). From Eq. 85 the free

energy of the additional quasiparticle is

εp − µ = εp − µ +∑p′

fp′,pδnp′ , (96)

(recall that fp,p′ = fp′,p). Both terms here are O(δ). The

second term describes the free energy of a quasiparticle due to

the other quasiparticles in the system (some sort of Hartree-like

term).

The term εp plays the part of the local energy of a quasi-

particle. For example, the gradient of εp is the force the system

exerts on the additional quasiparticle. When the quasiparticle

is added to the system, the system is inhomogeneous so that

δnp′ = δnp′(r). The system will react to this inhomogeneity

46

Page 47: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

by minimizing its free energy so that ∇rF = 0. However, only

the additional free energy due the added particle (Eq. 96) is

inhomogeneous, and has a non-zero gradient. Thus, the system

will exert a force

−∇rε = −∇r

∑p′

fp′,pδnp′(r) (97)

on the added quasiparticle resulting from interactions with other

quasiparticles.

4.2.1 Equilibrium Distribution of Quasiparticles at Finite T

εp also plays an important role in the finite-temperature prop-

erties of the system. If we write

E − E0 =∑

p

εpδnp +1

2

∑p,p′

fp′,pδnp′δnp (98)

Now suppose that∑

p |〈δnp〉| N , as needed for the expan-

sion above to be valid, so that

δnp = 〈δnp〉 + (δnp − 〈δnp〉) (99)

where the first term is O(δ), and the second O(δ2). Thus,

δnpδnp′ ≈ −〈δnp〉〈δnp′〉 + 〈δnp〉δnp′ + 〈δnp′〉δnp (100)

47

Page 48: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

We may use this to rewrite the energy of our interacting system

E − E0 ≈∑

p

εpδnp −1

2

∑p,p′

fp′,p〈δnp〉〈δnp′〉 +∑p,p′

fp′,p〈δnp〉δnp′

≈∑

p

εp +∑p′

fp′,p〈δnp′〉

δnp −1

2

∑p,p′

fp′,p〈δnp〉〈δnp′〉

≈∑

p

〈εp〉δnp −1

2

∑p,p′

fp′,p〈δnp〉〈δnp′〉 +O(δ4) (101)

At this point, we may repeat the arguments made earlier to de-

termine the fermion occupation probability for non-interacting

Fermions (the constant factor on the right hand-side has no

effect). We will obtain

np(T, µ) =1

1 + exp β(〈εp〉 − µ), (102)

or

δnp(T, µ) =1

1 + exp β(〈εp〉 − µ)− θ(pf − p) . (103)

However, at least for an isotropic system, this expression bears

closer investigation. Here, the molecular field (evaluated within

kBT of the Fermi surface)

〈εp − εp〉 =∑p′

fp′,p〈δnp′〉 (104)

48

Page 49: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

must be independent of the location of p on the Fermi sur-

face (and of course, spin), and is thus constant. To see this,

reconsider the Legendre polynomial expansion discussed earlier

〈εp − εp〉 =∑p′

fp′,p〈δnp′〉

∝∑l

∫d3pflPl(cos θ)〈δnp′〉

∝ f0

∫d3p〈δnp′〉 = 0

(105)

In going from the second to the third line above, we made use of

the isotropy of the system, so that 〈δnp′〉 is independent of the

angle θ. The evaluation in the third line, follows from particle

number conservation. Thus, to lowest order in δ

np(T, µ) =1

1 + exp β(εp − µ)+O(δ4) (106)

4.2.2 Local Equilibrium Distribution

Now suppose we introduce a local weak perturbation such as

the type discussed in Sec. 4.2 to an isotropic system at zero

temperature. Such a perturbation could be caused by, eg. a

49

Page 50: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

sound wave or a weak magnetic field (which you will explore in

your homework, to calculate the sound velocity and suscepti-

bility of a Landau Fermi liquid). This pertubation will cause a

small deviation of the equilibrium distribution function, leading

to a new ”local equilibrium” distribution

np = np(εp − µ) (107)

where the argument of the RHS indicates that this is the distri-

bution corresponding to the local energy discussed above. The

gradient of the local energy yields a force which tries to restore

the equilibrium distribution np(εp − µ) derived above). The

deviation from true equilibrium is

δnp = np − np (108)

= δnp +∂np(εp − µ)

∂εp(εp − εp) (109)

Using Eq. 96, we find

δnp = δnp −∂np

∂εp

∑p′

fp,p′δnp′ (110)

At zero temperature, the factor∂np∂εp

= −δ(εp−µ), so both δnp

and δnp are restricted to the fermi surface. Since the pertuba-

tion of interest is small, we may expand both δnp and δnp in

50

Page 51: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

a series of Legendre polynomials, and we will also split them

into symmetric and antisymmetric parts (as we did with fp,p′

previously). For example,

δnsp =∑l

δ(εp − µ)δnslPl (111)

If we make a similar expansion for the antisymmetric and sym-

metric parts of δnp, and substitute this back into Eq. 109, then

we find

δnal =

(1 +

F al

2l + 1

)δnal (112)

δnsl =

(1 +

F sl

2l + 1

)δnsl (113)

4.3 Effective Mass m∗ of Quasiparticles

This argument most closely follows that of AGD, and we will

follow their notation as closely as possible (without introducing

any new symbols). In particular, since an integration by parts

is necessary, we will use a momentum integral (as opposed to a

momentum sum) notation∑p

→ V

∫d3p

(2πh)3. (114)

51

Page 52: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

The net momentum of the volume V of quasiparticles is

Pqp = 2V

∫d3p

(2πh)3pnp net quasiparticle momentum (115)

which is also the momentum of the Fermi liquid. On the other

hand since the number of particles equals the number of quasi-

particles, the quasiparticle and particle currents must also be

equal

Jqp = Jp = 2V

∫d3p

(2πh)3vpnp net quasiparticle and particle current

(116)

or, since the momentum is just the particle mass times this

current

Pp = 2V m

∫d3p

(2πh)3vpnp net quasiparticle and particle current

(117)

where vp = ∇pεp, is the velocity of the quasiparticle. So∫d3p

(2πh)3pnp = m

∫d3p

(2πh)3∇pεpnp (118)

Now make an arbitrary change of np and recall that εp depends

upon np, so that

δεp = V∑σ′

∫d3p

(2πh)3fp,p′δnp′ . (119)

52

Page 53: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

For Eq. 118, this means that∫d3p

(2πh)3pδnp = m

∫d3p

(2πh)3∇pεpδnp (120)

+mV

∫d3p

(2πh)3

∑σ′

∫d3p′

(2πh)3∇p

(fp,p′δnp′

)np ,

or integrating by parts (and renaming p→ p′ in the last part),

we get∫d3p

(2πh)3

p

mδnp =

∫d3p

(2πh)3∇pεpδnp (121)

−V∑σ′

∫d3p′

(2πh)3

∫d3p

(2πh)3δnpfp,p′∇p′np′ ,

Then, since δnp is arbitrary, it must be that the integrands

themselves are equal

p

m= ∇pεp −

∑σ′

V

∫d3p′

(2πh)3fp,p′∇p′np′ (122)

The factor ∇p′np′ = −p′

p′ δ(p′− pF ). The integral may be eval-

uated by taking advantage of the system isotropy, and setting

p parallel to the z-axis, since we mostly interested in the prop-

erties of the system on the Fermi surface we take p = pF , let θ

be the angle between p (or the z-axis) and p′, and finally note

53

Page 54: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

that on the Fermi surface∣∣∣∇pεp|p=pF

∣∣∣ = vF = pF/m∗. Thus,

pFm

=pFm∗

+∑σ′

∫p′2dpdΩ

(2πh)3fpσ,p′σ′

p′

p′δ(p′ − pF ) (123)

However, since both p and p′ are restricted to the Fermi surface

p′

p′ = cos θ, and evaluating the integral over p, we get

1

m=

1

m∗+V pF

2

∑σ,σ′

∫dΩ

(2πh)3fpσ,p′σ′ cos θ , (124)

where the additional factor of 12 compensates for the additional

spin sum. If we now sum over both spins, σ and σ′, only the

symmetric part of f survives (the sum yields 4f s), so

1

m=

1

m∗+

4πV pF(2πh)3

∫d (cos θ) f s(θ) cos θ , (125)

We now expand f in a Legendre polynomial series

fα(θ) =∑l

fαl Pl(cos θ) , (126)

and recall that P0(x) = 1, P1(x) = x, .... that∫ 1

−1

dxPn(x)Pm(x)dx =2

2n + 1δnm (127)

and finally that

D(0)fαl =V m∗pF

π2h3 fαl = F αl , (128)

54

Page 55: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

we find that1

m=

1

m∗+

F s1

3m∗, (129)

or m∗/m = 1 + F s1/3.

Quantity Fermi Liquid Fermi Liquid/Fermi Gas

Specific Heat Cv = m∗pF3h3

k2BT

CV

CV 0= m∗

m= 1 + F s

1 /3

Compressibility κκ0

=1+F s

0

1+F s0 /3

Sound Velocity c2 =p2F

3mm∗ (1 + F s0 )

(cc0

)2

=1+F s

0

1+F s1 /3

Spin Susceptibility χ = m∗pFπ2h3

β2

1+Fa0

χχ0

=1+F s

1 /3

1+Fa0

Table 1: Fermi Liquid relations between the Landau parameters Fαn and some exper-

imentally measurable quantities. For the latter, a zero subscript indicates the value

for the non-interacting Fermi gas.

The effective mass cannot be experimentally measured di-

rectly; however, it appears in many physically relevant measur-

able quantities, including the specific heat

CV =

(∂E/V

∂T

)V N

=1

V

∂T

∑p

εpnp. (130)

To lowest order in δ, we may neglect fp,p′ in both εp and np,

so

CV =1

V

∑p

εp∂np

∂T. (131)

Recall that the density of states D(E) =∑

p δ(E − εp), and

making the same assumption that we made for the non-interacting

55

Page 56: Chapter 6: The Fermi Liquid - LSUjarrell/COURSES/SOLID_STATE/Chap6/chap6.pdf · 1 introduction: The Electronic Fermi Liquid As we have seen, the electronic and lattice degrees of

system, that ∂µ∂T is negligible, we get,

CV =1

V

∫dεD(ε)ε

∂T

1

exp β(ε− µ) + 1. (132)

This integral is identical to the one we had to evaluate for the

non-interacting system, and yields the result

CV =π2

3Vk2BTD(EF )

=k2BTm

∗pF

3h3 . (133)

Thus, measuring the electronic contribution to the specific heat

CV yields information about the effective mass m∗, and hence

F s1 . Other measurements are related to some of the remaining

Landau parameters, as summarized in table 1.

References

[1] https://arxiv.org/pdf/hep-ph/0407258v2.pdf

[2] https://arxiv.org/pdf/1411.4372.pdf

56